Fabricated Cu–Ag Nanoparticle-Embedded ... - ACS Publications

Jan 5, 2017 - In Situ Fabricated Cu−Ag Nanoparticle-Embedded Polymer Thin Film as an Efficient Broad Spectrum SERS Substrate .... system and INCA an...
0 downloads 12 Views 8MB Size
Article pubs.acs.org/JPCC

In Situ Fabricated Cu−Ag Nanoparticle-Embedded Polymer Thin Film as an Efficient Broad Spectrum SERS Substrate V. Kesava Rao, Pankaj Ghildiyal, and T. P. Radhakrishnan* School of Chemistry, University of Hyderabad, Hyderabad - 500 046, India S Supporting Information *

ABSTRACT: Thin films of polymer−metal nanocomposites can serve as efficient surface-enhanced Raman scattering (SERS) substrates; it would be highly advantageous if it can be fabricated by a cost-effective and facile protocol, and high enhancement factors (EF) can be realized for a range of excitation wavelengths, enabling sensitive detection of different analytes. Poly(vinyl alcohol) (PVA) thin films with embedded Cu−Ag nanoparticles are fabricated through a simple in situ procedure via stepwise formation of CuO−PVA and Cu−PVA; the fabrication is monitored through the different stages, by spectroscopy, microscopy, electron diffraction, and elemental analysis. Choice of the Cu−Ag combination is dictated by its broad plasmonic extinction, favorable cost factor, and chemical stability of the nanocomposite; the hydrogel character of PVA facilitates analyte absorption and contact with the nanoparticles, leading to efficient SERS. Cu−Ag−PVA thin film with an optimal composition is shown to display plasmonic extinction over an appreciable part of the visible range and to provide EF of the order of 107−108 for the analytes, rhodamine 6G and methylene blue, at three different excitation wavelengths; the EF values are significantly higher than those reported earlier, with most of the Cu−Ag based SERS substrates. The novel polymer−bimetal nanocomposite thin film is an efficient SERS substrate providing subpicomolar limits of detection.



INTRODUCTION

Cu and Ag form a binary alloy system with a eutectic at 42 atom % Cu.7,8 The relatively lower surface energy and larger atomic size of Ag induce phase separation in the alloy, and core−shell structures emerge in nanoparticles within some size limits.9 Formation of a Ag shell on Cu nanoparticles imparts stability against oxidation, as has been shown in the case of airstable dispersions for inkjet printing.10 Treatment of Cu with Ag+ ions is a facile route to deposit Ag on Cu through galvanic displacement; this approach has been adopted in a variety of ways, to fabricate SERS substrates based on the Cu−Ag bimetallic system. The most common protocol involves the formation of Ag nano/microstructures with a variety of morphologies, on the surface of Cu in the form of grid,11 foil,12−14 plate,15,16 membrane,17 layer deposits,18−20 and powder;21 Ag nanoparticles have also been formed on Cu nanowalls in a microfluidic device.22 Cu−Ag nanoparticles,23,24 dendrites,25,26 and core−shell structures4,27 formed in a fluid phase have also been developed for SERS application. Bimetallic nanoparticles embedded in a suitable thin film matrix that can efficiently absorb analyte molecules should prove to be a versatile SERS substrate. An optimal substrate that provides high EF for different analyte molecules employing different excitation wavelengths remains an important goal.

Surface enhanced Raman scattering (SERS) is a facile and powerful analytical tool for the sensitive detection of molecules, thanks to the combination of the unique Raman spectral signature of molecules with extremely large signal enhancements. The latter is often achieved through the interaction of the molecules with metal nanoparticles and nanostructures, and the field associated with their localized surface plasmon resonance (LSPR) excitation. Metals such as gold, silver, and copper with LSPR extinction in the 400−700 nm range are suitable for visible light excitation; even though silver and copper provide obvious cost advantages, the relatively higher reactivity is a cause for concern in practical applications. The large extinction cross section associated with silver nanoparticles has been exploited extensively;1 copper nanoparticles and nanostructures have also been shown to be efficient substrates for SERS.2,3 As the LSPR extinction spectra of Ag and Cu nanoparticles peak at ∼400 and ∼600 nm, respectively, an optimal combination of the two should provide local field enhancement over a significant part of the visible spectrum, and lead to efficient SERS with different excitation wavelengths;4 strong dependence of the SERS enhancement factor (EF) on the excitation wavelength is well established.5,6 A Cu−Ag nanostructure with appropriate composition would serve as an efficient SERS substrate for different analyte molecules, through resonance effects; practical application would be facilitated if the substrate is stable and in the form of a solid thin film. © XXXX American Chemical Society

Received: October 10, 2016 Revised: December 3, 2016

A

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 1. Schematic showing the fabrication of Cu−PVA thin film and its conversion to Cu−Ag−PVA thin film.

In situ formation of metal28−30 and bimetallic31,32 nanoparticles inside polymer thin films is a facile route to the fabrication of efficient SERS substrates.33 Ag, Au, Pd, and Pt nanoparticles have been fabricated inside a poly(vinyl alcohol) (PVA) thin film by mild thermal annealing using the polymer itself as the reducing and stabilizing agent.30 As this protocol is ineffective in the case of Cu in view of its relatively lower reduction potential, we have now optimized a new methodology for the formation of Cu nanoparticle-embedded PVA thin film (Cu−PVA) through the intermediate formation of CuO− PVA. Brief treatment of the Cu−PVA film with AgNO3 solution of varying concentration under ambient conditions led to the formation of Cu−Ag−PVA thin films with different Cu−Ag compositions. It is shown that the Cu−Ag−PVA thin films provide SERS EF of the order of 107−108 (higher than that reported earlier, with most of the Cu−Ag based substrates) for two different analyte molecules, with three different laser excitation wavelengths; improved efficiency over the corresponding single metal systems, Cu−PVA and Ag−PVA with similar metal loading is also demonstrated.

dish (150 mm diameter, 15 mm high) together with 3 mL of 80% hydrazine hydrate (Finar Chemicals, India) taken in an open Petri dish (50 mm diameter, 12 mm high). The whole system was kept in a convection oven at 160 °C for 20 min (Figure 1), to transform CuO−PVA to Cu−PVA. Fabrication of Cu−Ag−PVA. Cu−PVA film (x = 0.25) coated on the substrate was used for the synthesis of the bimetallic nanoparticles. 0.5 mL of an aqueous solution of AgNO3 (Aldrich, purity ≥99.0%) with concentrations ranging from 1 to 12 mM was spread uniformly on the Cu−PVA film and kept for 10 min under ambient temperature (28 ± 1 °C) inside a closed Petri dish to avoid evaporation of the solution. The film was then washed with ∼1 mL of water and dried under ambient atmosphere; even though PVA is soluble in water, the Cu−Ag−PVA film is insoluble, most likely due to partial cross-linking that occurs during the thermal treatment with metal salts.34 Fabrication of Ag−PVA. Silver nitrate (160 mg, Aldrich, purity ≥99.0%) dissolved in 0.8 mL of water was mixed with a solution of 0.10 g of PVA in 2 mL of water (Ag/PVA weight ratio = 1.0). The solution mixture was spin-coated on different substrates as described earlier, and heated at 130 °C for 60 min; in an alternate approach, the film was exposed to hydrazine vapor at 28 ± 1 °C for 10 min. Characterization of the Polymer−Metal Nanocomposite Thin Films. Film thickness was measured using an Ambios Technology XP-1 Profilometer (Figure S1). The elemental composition of the films was analyzed using a Varian Model Liberty Series Inductively Coupled Plasma-Optical Emission Spectrometer (ICP-OES); the sample for the analysis was prepared by dissolving the film in 100 mL of 69% nitric acid. Electronic extinction spectra of the thin films coated on a quartz substrate were recorded on a Varian Model Cary 100 UV−visible spectrometer. Transmission electron microscopy (TEM) was carried out on an FEI TECNAI G2 S-Twin TEM at an accelerating voltage of 200 kV. Field emission scanning electron microscope (FE-SEM) imaging with energy dispersive X-ray spectroscopy (EDXS) was carried out on a Carl Zeiss model Ultra 55 microscope; EDX spectra and maps were recorded using an Oxford Instruments X-MaxN SDD (50 mm2)



EXPERIMENTAL SECTION Fabrication of CuO−PVA. The required weight of Cu(NO3)2·3H2O (Aldrich, purity ≥99.0%) dissolved in 1.6 mL of water was mixed with 0.20 g of PVA (Aldrich, average molecular weight = 85−146 kDa, % hydrolysis = 99+) dissolved in 4.0 mL of water to prepare different compositions designated using the Cu/PVA weight ratios, x (for example, 76 mg of Cu(NO3)2·3H2O gives x = 0.10); all solutions were prepared using Millipore Milli-Q water (resistivity = 18.2 MΩ cm). Glass/quartz/Si wafer substrates used for coating the films were cleaned by washing with soap and water, keeping in nitric acid followed by washing in water, sonication in isopropyl alcohol, and finally drying in a hot air oven. The solution mixture was spin-coated on the substrate using a Laurell Technologies Corporation Model WS-650HZ-23NPP/LITE Photoresist Spinner at 500 rpm for 10 s, followed by 6000 rpm for 10 s. The film was heated in a convection oven at 130 °C for 60 min. Fabrication of Cu−PVA. CuO−PVA film (x = 0.10, 0.25, and 0.50) coated on the substrate was kept in a closed Petri B

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 2. Electronic absorption/extinction spectra of (a) Cu(NO3)2−PVA and CuO−PVA, and (b) Cu−PVA thin films with different weight ratios (x) of Cu/PVA.

NSERS) × (ISERS/IBulk), the intensities of different vibrational peaks were chosen, based on the excitation wavelength and analyte (Figure S12, Tables S6−S8). As NSERS includes all the probe molecules in the focal volume and not just a monolayer on the nanoparticles, the EF values we report are the lower bounds.

system and INCA analysis software. The preparation of the samples for TEM and FE-SEM imaging was carried out as follows: a glass substrate was first coated with a few drops of a solution of polystyrene (Aldrich, average molecular weight = 280 kDa) in toluene (1 g in 7 mL) by spinning at 1000 rpm for 10 s and dried at 90 °C for 15 min; this substrate was used for the fabrication of the various films as described earlier, and the multilayer film obtained was peeled off the glass substrate, placed on a 200 mesh copper grid, and then dipped in toluene, whereupon the polystyrene layer dissolves and the polymer− metal nanocomposite film sticks to the grid. SERS Experiments. Experiments were conducted using Cu−Ag−PVA (with different Ag contents) coated on silicon wafer as the SERS substrate; control experiments were carried out using Cu−PVA and Ag−PVA films. Methanol solutions of rhodamine 6G (R6G, Aldrich) (20 μL, 8.6 μM) and methylene blue (MB, Aldrich) (20 μL, 6.3 μM) were spread uniformly on the substrate and dried under ambient atmosphere (concentrations determined using measured absorbance and reported extinction coefficient). A WITec model Alpha 300 R Raman microscope (with AFM) was used for recording the Raman spectra, with 0.5 s integration time and 10 accumulations, through a 20× aperture (NA = 0.4). Three different line lasers (488, 633, and 785 nm) were used as the excitation source; the power at the focus (spot diameter) was 4.8 mW (1.49 μm), 13.4 mW (1.93 μm), and 10.5 mW (2.39 μm) respectively. The laser intensity was maintained constant in all measurements; 100 μm (for 488 nm) and 50 μm (for 633 and 785 nm) detecting fibers were used to collect the spectra. Raman spectra for the bulk material used as a reference were recorded using a small R6G and MB microcrystal placed on the Si wafer. The number of molecules in the reference (NBulk) was determined using the focal volume of the laser spot incident on the microcrystal and the density of the material; the number of molecules in the sample (NSERS) was estimated by considering the total number of molecules spread on the substrate, assuming a homogeneous distribution across the thin film and the fraction falling within the laser spot area. Raman spectra recorded with the pure PVA film as well as Cu−Ag− PVA film showed a negligible background, but the spectra of the films with the analyte had a broad background due to fluorescence (Figure S14); the spectra presented are corrected for this background, and used for analyzing the spectral line intensities. To estimate the enhancement factor, EF = (NBulk/



RESULTS AND DISCUSSION Among the metals fabricated and studied as nanoparticles and nanostructures, Au and Ag dominate significantly above all others. Even though relatively less, nanoparticles of the remaining coinage metal, Cu, have also been studied extensively.35 The lower reduction potential of Cu+/Cu and Cu2+/Cu compared to Ag+/Ag or Au3+/Au necessitates stronger reducing agents for the formation of Cu; the associated ease of oxidation of Cu makes the stabilization of Cu nanoparticles more difficult. In situ reduction of the metal precursor ion by the PVA itself acting as the reducing agent, that we have successfully employed30 in the case of metals like Ag, Au, Pd, Hg, and Pt, is not possible in the case of Cu. However, if formed inside the polymer film, the Cu nanoparticles are likely to be relatively stable against atmospheric oxidation. Therefore, we have developed a twostep protocol for the fabrication of Cu−PVA thin films (Figure 1). Spin-coated thin films of Cu(NO3)2−PVA with three different Cu/PVA weight ratios (x) were heated at 130 °C for 1 h to yield CuO−PVA thin films, the oxide being formed by thermal decomposition of the nitrate salt. Interestingly, heating pure Cu(NO3)2·3H2O powder at 130 °C under an ambient atmosphere produces only a color change from blue to cyan (indicating the formation of basic copper nitrate), whereas heating a solid mixture of Cu(NO3)2·3H2O and PVA produces a black solid, suggesting the formation of CuO (Figures S2 and S3). Thermal decomposition appears to be facilitated by the coordination of the alcoholic groups on PVA to Cu2+; absence of the characteristic red color of Cu2O rules out the possibility of reduction of Cu2+ during this process. Formation of CuO nanoparticles is demonstrated by the development of the characteristic band absorption starting at 300−550 nm, which increases with x (Figure 2a). Tauc plots indicate direct band gaps in the range 3.36−4.01 eV for the different compositions (Figure S4); these values are consistent with those reported for CuO nanoparticles.36,37 TEM images of the films reveal small particles, 1−10 nm in size depending on x (Figure 3). The high C

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 3. TEM images (scale = 20 nm) of CuO−PVA thin films with weight ratios (x) of (a) 0.10, (b) 0.25, and (c) 0.50; the corresponding magnified images (scale = 5 nm) are shown in parts d−f; the electron diffraction pattern and lattice spacing are indicated in the insets in parts c and f, respectively.

Figure 4. TEM images (scale = 20 nm) of Cu−PVA thin films with weight ratios (x) of (a) 0.10, (b) 0.25, and (c) 0.50; the corresponding magnified images (scale = 5 nm) are shown in parts d−f; the electron diffraction pattern and lattice spacing are indicated in the insets in parts a−c and e, respectively.

resolution TEM image of the film with x = 0.5 shows the expected lattice spacing; the electron diffraction pattern can be indexed to monoclinic CuO (JCPDS file no: 89-2531). In the second stage, CuO nanoparticles formed inside the PVA film were reduced in situ using hydrazine vapor, by maintaining the film at 160 °C for 20 min. The reduction is reflected clearly in the extinction spectrum in Figure 2b which shows the emergence of the LSPR peak at ∼580−590 nm. The peak intensity is directly related to the Cu/PVA ratio; the

scattering contribution increases at higher values of x. TEM images of the Cu−PVA films show the presence of Cu nanoparticles with sizes in the range 10−30 nm (Figure 4). The identity of Cu is proved by the high resolution TEM image showing the lattice planes of Cu, as well as the electron diffraction pattern (Table S4); XPS data (Figure S9) are consistent with the formation of Cu with signs of slight oxidation. Cu nanoparticles can indeed be generated by the direct treatment of Cu(NO3)2−PVA with hydrazine; however, D

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C the quality of the Cu−PVA films obtained is poor. Cu−PVA films obtained in the two-step route via thermal processing are smooth (AFM measured surface roughness is ∼0.5 nm (Figure S5; Table S3)), optically clear, and uniform (verified by the LSPR spectra recorded at several points on the film); the LSPR extinction spectrum is stable for several hours but shows small variation over a few days (Figure S10). It may be noted that the formation of Cu−Ag particles imparts high stability to the system as discussed below. Cu−Ag nanoparticles were formed in situ by treating the Cu−PVA thin films with AgNO3 solution having different concentrations (c), for 10 min in each case; as the film with x = 0.25 showed the clear plasmon peak due to Cu with relatively low scattering, it was chosen for the incorporation of Ag and used in all of the studies discussed further. Galvanic replacement of Cu by Ag is revealed by the emergence of the LSPR extinction due to Ag with λmax at ∼400−450 nm (Figure 5). The amplitude as well as width of the extinction and hence

mapping shown in Figure 8; it is also seen that the two elements are distributed fairly evenly throughout the particle. A parallel trend in the Cu−Ag composition is observed in the ICP analysis data obtained using bulk film samples of Cu−Ag−PVA (Tables S1 and S2); differences in the absolute values from EDXS may be attributed to the fact that the latter probes primarily the surface region of the thin films. XPS data show the presence of Cu (along with Cu2+, most likely due to CuO) and Ag in the Cu−Ag−PVA film (Figure S9). The Cu−Ag−PVA films coated on glass substrates are ∼300 nm thick (Figure S1). As discussed above, these films exhibit strong LSPR extinction over a broad range in the visible regime, due to the embedded bimetallic nanoparticles; the hydrogel character of PVA facilitates efficient swelling and absorption of analyte solutions. These characteristics are expected to make these nanocomposite thin films cheap, easily fabricated, and efficient substrates for the sensitive SERS detection of analytes using different excitation wavelengths. In order to assess the efficiency of the Cu−Ag−PVA thin films as SERS substrates for broad spectrum applications, Raman spectra were recorded for methanol solutions of rhodamine 6G (R6G) and methylene blue (MB) which show strong visible light absorption at 460− 560 and 580−680 nm, respectively (Figure S12); three excitation wavelengths, 488, 633, and 785 nm, were employed in the measurements. Cu−Ag−PVA prepared from Cu−PVA (x = 0.25) treated with different concentrations of AgNO3 showed the maximum EF values in the SERS experiments compared to the films prepared from Cu−PVA with other x values, and hence were used in the Raman studies; control experiments were carried out using Cu−PVA and Ag−PVA films as the substrate. Raman spectra recorded for R6G (λexc = 488 nm) and MB (λexc = 633 nm) are presented in parts a and b of Figure 9, respectively. The spectrum recorded for the reference microcrystal is shown in each case; the peaks are hardly visible but can be discerned upon expanding the scale. The spectrum of R6G recorded with Cu−PVA as the substrate shows slightly improved intensities. Incorporation of Ag by treatment with AgNO3 solution of increasing concentrations led to a significant increase in the spectral intensity with the maximum observed with c = 12 mM; aggregation of particles in the films fabricated with c > 12 mM possibly leads to decreasing SERS efficiency. The spectrum recorded on Ag−PVA (with a similar content of Ag (Table S1)) is significantly weaker, with intensity comparable to the Cu−Ag−PVA film formed by treatment with 1 mM AgNO3; a comparison of the spectra obtained on Cu−Ag−PVA (c = 12 mM) film and Ag−PVA film having similar size nanoparticles (Figure S8) shows the former to be significantly stronger, revealing the critical role of the bimetallic composition. Parallel trends are seen in the experiments with MB. Spectra were recorded with the other two laser excitations in each case. In the case of R6G, 785 nm excitation produces relatively weak spectral lines (Figure S15); the spectral intensities are very high with the excitations at 633 and 488 nm, the low energy vibrations being stronger with the former and the high energy vibrations with the latter. With MB, 488 nm excitation produces weaker spectra than 633 and 785 nm excitations (Figure S16); in the latter two cases once again, the lower energy vibrations dominate with the higher excitation wavelength and vice versa. These observations of selective enhancement of vibrational modes related to the excitation wavelength are similar to the earlier observations on R6G on Ag colloids, and MB on random array of gold nanoparticles

Figure 5. Extinction spectra of Cu−PVA (x = 0.25) and Cu−Ag−PVA thin films fabricated by treating the Cu−PVA film with different concentrations (c) of AgNO3 solution.

the coverage over the 350−700 nm increase with the concentration (c) up to 12 mM; the spectrum changes very little beyond this point, indicating saturation (in fact, the 450 nm peak decreases marginally and more scattering is observed at higher wavelengths (Figure S6)). The nanocomposite films are extremely stable, as shown by the reproducibility of the extinction spectrum even after several days or on exposure to O2 gas (Figures S10 and S11), the stability being attributable to the incorporation of Ag with a higher reduction potential than Cu; exposure to reactive gases like H2S however does affect the film (Figure S11). TEM images of the Cu−Ag−PVA thin films are collected in Figures 6 and 7; the electron diffraction patterns are shown in the insets of the former, and the latter with higher magnification show the lattice planes. Nearly spherical particles are observed in all cases, with the average particle size increasing from (a) to (e); the spectral changes are consistent with this. The lattice spacing of 2.34 Å can be attributed to the (111) planes of Ag; the spacing of 2.08−2.10 Å could be due either to the (200) planes of Ag or the (111) planes of Cu (Table S5).25 EDXS data clearly show the presence of Cu and Ag, with the atom % of Ag in the films fabricated with c = 1−12 mM, increasing from 29 to 81, respectively (Table S2). The Cu−Ag composition variation can be visualized from the FESEM images combined with EDXS E

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 6. TEM images (scale = 50 nm) of Cu−Ag−PVA thin films fabricated by treating Cu−PVA (x = 0.25) with different concentrations (c, mM) of AgNO3 solution: (a) 1, (b) 3, (c) 6, (d) 9, and (e) 12. Electron diffraction patterns are indicated in the insets.

Figure 7. TEM images (scale = 5 nm) of Cu−Ag−PVA thin films fabricated by treating Cu−PVA (x = 0.25) with different concentrations (c, mM) of AgNO3 solution: (a) 1, (b) 3, (c) 6, (d) 9, and (e) 12; lattice spacing values are indicated and discussed in the text.

covered with silica.6,38 The EF estimated from the various experiments for the analytes R6G and MB are collected in Tables 1 and 2, respectively; appreciable EF values in the 107− 108 range are observed in most of the cases. It is seen also that the excitation wavelength of 488 nm provides the highest EF with the different Cu−Ag−PVA films for R6G, whereas 633 nm is the best suited for MB; these observations are consistent with the absorption profile of the two analyte molecules, and point to potential resonance effects in the Raman scattering response. The Cu−Ag nanoparticle substrate provides not only strong plasmonic fields over a wide range of wavelengths facilitating

excitation by different lasers but also efficient enhancement due to the interfaces that form between the bimetallic particles. Comparison with the Cu−Ag based substrates reported earlier shows that the SERS EF values obtained with Cu−Ag−PVA are significantly higher in most of the cases (Table S9). We have examined the linearity of the SERS response with respect to the analyte concentration and the limit of detection (LOD). Raman spectra recorded for different concentrations of R6G (λexc = 488 nm) and MB (λexc = 633 nm) spread on the Cu−Ag−PVA film substrate (fabricated with c = 12 mM) are shown in Figure 10a and b. Plots of the intensities of the 1645 F

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 8. FESEM images combined with EDXS area mapping, of (a) Cu−PVA (x = 0.25) and Cu−Ag−PVA thin films formed by treating the Cu− PVA with different concentrations (c, mM) of AgNO3 solution: (b) 1, (c) 3, (d) 6, (e) 9, and (f) 12; scale bar = 100 nm. The EDXS mapping shows Cu (red) and Ag (green) regions in a selected particle.

Figure 9. Raman spectra of (a) R6G (λexc = 488 nm) and (b) MB (λexc = 633 nm) on Cu−Ag−PVA thin films fabricated by treatment of Cu−PVA (x = 0.25) with different concentrations (c) of AgNO3; spectra on Cu−PVA and Ag−PVA thin films and for the reference microcrystal (R6G or MB) are also shown in each case.

Table 1. Enhancement Factor for the Raman Spectra of R6G Recorded (with Different λexc) on Cu−Ag−PVA Thin Films Fabricated by Treatment of Cu−PVA (x = 0.25) with Different Concentrations (c) of AgNO3; Values for Cu−PVA and Ag−PVA are Also Showna

Table 2. Enhancement Factor for the Raman Spectra of MB Recorded (with Different λexc) on Cu−Ag−PVA Thin Films Fabricated by Treatment of Cu−PVA (x = 0.25) with Different Concentrations (c) of AgNO3; Values for Cu−PVA and Ag−PVA are Also Showna

EF (×108) for λexc substrate Cu−PVA Cu−Ag−PVA [c (mM)]

Ag−PVA

1 3 6 9 12

488 nm

633 nm

785 nm

0.06 0.41 0.77 1.44 1.77 2.27 0.86

0.01 0.19 0.26 0.38 1.31 1.60 0.09

0.01 0.02 0.02 0.03 0.03 0.002 ∼0.00

EF (×108) for λexc substrate Cu−PVA Cu−Ag−PVA [c (mM)]

Ag−PVA

1 3 6 9 12

488 nm

633 nm

785 nm

0.03 0.18 0.17 0.19 0.08 0.06 ∼0.00

0.05 1.37 2.09 2.37 3.04 3.49 0.22

0.05 0.13 0.18 0.20 0.34 0.61 ∼0.00

a

a

cm−1 peak of R6G and 1620 cm−1 peak of MB versus the analyte solution concentration show clear linear variation (Figure 10c and d); the very small standard deviations of the measured spectral intensities attest to the uniformity of the

substrate as well as the analyte distribution on it (Figure S13). The LOD (defined as 3σblank/m, where σblank is the standard deviation for blank measurements and m is the slope of the intensity−concentration plot)39,40 estimated are 0.66 and 0.32

The intensity of the strongest vibrational peak was chosen to calculate the EF in each case.

The intensity of the strongest vibrational peak was chosen to calculate the EF in each case.

G

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 10. Raman spectra for varying concentrations of (a) R6G (λexc = 488 nm) and (b) MB (λexc = 633 nm) on Cu−Ag−PVA thin film fabricated by treatment of Cu−PVA (x = 0.25) with 12 mM AgNO3 solution. Plot of the intensity of (c) 1645 cm−1 of the R6G spectra and (d) the 1620 cm−1 peak of the MB spectra versus concentration; standard deviations in the data and the best fit line are shown.



pmol for R6G and MB, respectively. These observations point to the high sensitivities of detection that can be achieved using the Cu−Ag−PVA thin films as SERS substrates.



Corresponding Author

*E-mail: [email protected]. Fax: 91-40-2301-2460. Phone: 9140-2313-4827.

CONCLUSION

ORCID

T. P. Radhakrishnan: 0000-0002-0318-4461

The present study demonstrates a facile route to the formation of stable bimetal−polymer nanocomposite thin films using the cost-effective choice of noble metals, Ag and Cu, serving as an efficient broad spectrum SERS substrate. Thin film with an optimal composition is shown to produce high enhancement factors for different analytes at varying excitation wavelengths. The general concept demonstrated can be implemented with other nanocomposite thin films to realize high SERS responses over broad spectral ranges.



AUTHOR INFORMATION

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support from the Department of Science and Technology and the University Grants Commission (BSR and UPE programs), New Delhi, and infrastructure support from the Central Facility for Nanotechnology and the School of Chemistry at the University of Hyderabad are acknowledged with gratitude. We thank Mr. Muvva Durgaprasad and Mr. D. Sunil for help with the TEM and FESEM imaging and Mr. S. S. Deo (National Chemical Laboratory, Pune) for the XPS measurements. V.K.R. thanks the UGC, New Delhi, for a senior research fellowship.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.6b10238. Details of thickness, elemental composition analysis, film morphology, extinction spectra, electron diffraction analysis, and XPS of Cu−Ag−PVA films; calculation of the SERS enhancement factors for films with varying composition, and different analytes and excitation wavelengths, along with estimation of the limit of detection (PDF)



REFERENCES

(1) Schlücker, S. Surface-Enhanced Raman Spectroscopy: Concepts and Chemical Applications. Angew. Chem., Int. Ed. 2014, 53, 4756− 4795. (2) Pereira, A. J.; Gomes, J. P.; Lenz, G. F.; Schneider, R.; Chaker, J. A.; de Souza, P. E. N.; Felix, J. F. Facile Shape-Controlled Fabrication of Copper Nanostructures on Borophosphate Glasses: Synthesis, Characterization, and Their Highly Sensitive Surface-Enhanced Raman H

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C Scattering (SERS) Properties. J. Phys. Chem. C 2016, 120, 12265− 12272. (3) Tan, Y.; Gu, J.; Xu, W.; Chen, Z.; Liu, D.; Liu, Q.; Zhang, D. Reduction of CuO Butterfly Wing Scales Generates Cu SERS Substrates for DNA Base Detection. ACS Appl. Mater. Interfaces 2013, 5, 9878−9882. (4) Lee, J.; Chen, D.; Li, X.; Yoo, S.; Bottomley, L. A.; El-Sayed, M. A.; Park, S.; Liu, M. Well-Organized Raspberry-Like Ag@Cu Bimetal Nanoparticles for Highly Reliable and Reproducible Surface-Enhanced Raman Scattering. Nanoscale 2013, 5, 11620−11624. (5) McFarland, A. D.; Young, M. A.; Dieringer, J. A.; Van Duyne, R. P. Wavelength-Scanned Surface-Enhanced Raman Excitation Spectroscopy. J. Phys. Chem. B 2005, 109, 11279−11285. (6) Á lvarez-Puebla, R. A. Effects of the Excitation Wavelength on the SERS Spectrum. J. Phys. Chem. Lett. 2012, 3, 857−866. (7) Sopousek, J.; Zobac, O.; Bursik, J.; Roupcova, P.; Vykoukal, V.; Broz, P.; Pinkas, J.; Vrestal, J. Heat-Induced Spinodal Decomposition of Ag−Cu Nanoparticles. Phys. Chem. Chem. Phys. 2015, 17, 28277− 28285. (8) Chen, H.; Zuo, J. Structure and Phase Separation of Ag−Cu Alloy Thin Films. Acta Mater. 2007, 55, 1617−1628. (9) Martinez De La Hoz, J. M.; Tovar, R. C.; Balbuena, P. B. Size Effect on the Stability of Cu−Ag Nanoalloys. Mol. Simul. 2009, 35, 785−794. (10) Grouchko, M.; Kamyshny, A.; Magdassi, S. Formation of AirStable Copper−Silver Core−Shell Nanoparticles for Inkjet Printing. J. Mater. Chem. 2009, 19, 3057−3062. (11) Guo, T.; Li, J.; Sun, X.; Sakka, Y. Improved Galvanic Replacement Growth of Ag Microstructures on Cu Micro-Grid for Enhanced SERS Detection of Organic Molecules. Mater. Sci. Eng., C 2016, 61, 97−104. (12) Jiang, X.; Lai, Y.; Yang, M.; Yang, H.; Jiang, W.; Zhan, J. Silver Nanoparticle Aggregates on Copper Foil for Reliable Quantitative SERS Analysis of Polycyclic Aromatic Hydrocarbons with a Portable Raman Spectrometer. Analyst 2012, 137, 3995−4000. (13) Mabbott, S.; Larmour, I. A.; Vishnyakov, V.; Xu, Y.; Graham, D.; Goodacre, R. The Optimisation of Facile Substrates for Surface Enhanced Raman Scattering through Galvanic Replacement of Silver onto Copper. Analyst 2012, 137, 2791−2798. (14) Hao, J.; Xu, Z.; Han, M.; Xu, S.; Meng, X. Surface-Enhanced Raman Scattering Analysis of Perchlorate Using Silver Nanofilms Deposited on Copper Foils. Colloids Surf., A 2010, 366, 163−169. (15) Gao, T.; Wang, Y.; Wang, K.; Zhang, X.; Dui, J.; Li, G.; Lou, S.; Zhou, S. Controlled Synthesis of Homogeneous Ag NanosheetAssembled Film for Effective SERS Substrate. ACS Appl. Mater. Interfaces 2013, 5, 7308−7314. (16) Pergolese, B.; Muniz-Miranda, M.; Bigotto, A. Surface-Enhanced Raman Scattering Investigation of the Adsorption of 2-Mercaptobenzoxazole on Smooth Copper Surfaces Doped with Silver Colloidal Nanoparticles. J. Phys. Chem. B 2006, 110, 9241−9245. (17) Chen, H.; Luo, J.; Zeng, T.; Jiang, L.; Sun, Y.; Jiao, Z.; Jin, Y.; Sun, X. Investigation of the Synthesis, SERS Performance and Application in Glucose Sensing of Hierarchical 3D Silver Nanostructures. New J. Chem. 2014, 38, 3907−3916. (18) Zhang, Q.; Chen, Y.; Guo, Z.; Liu, H.; Wang, D.; Huang, X. Bioinspired Multifunctional Hetero-Hierarchical Micro/Nanostructure Tetragonal Array with Self-Cleaning, Anticorrosion, and Concentrators for the SERS Detection. ACS Appl. Mater. Interfaces 2013, 5, 10633−10642. (19) Ke, X.; Lu, B.; Hao, J.; Zhang, J.; Qiao, H.; Zhang, Z.; Xing, C.; Yang, W.; Zhang, B.; Tang, J. Facile Fabrication of SERS Arrays through Galvanic Replacement of Silver onto Electrochemically Deposited Copper Micropatterns. ChemPhysChem 2012, 13, 3786− 3789. (20) Jie, Z.; Pengyue, Z.; Yimin, D.; Xiaolei, Z.; Jiamin, Q.; Yong, Z. Ag-Cu Nanoparticles Encaptured by Graphene with Magnetron Sputtering and CVD for Surface-Enhanced Raman Scattering. Plasmonics 2016, 11, 1495.

(21) Kim, K.; Lee, H. S. Effect of Ag and Au Nanoparticles on the SERS of 4-Aminobenzenethiol Assembled on Powdered Copper. J. Phys. Chem. B 2005, 109, 18929−18934. (22) Parisi, J.; Su, L.; Lei, Y. In Situ Synthesis of Silver Nanoparticle Decorated Vertical Nanowalls in a Microfluidic Device for Ultrasensitive In-Channel SERS Sensing. Lab Chip 2013, 13, 1501−1508. (23) Zhang, X.; Xu, S.; Jiang, S.; Wang, J.; Wei, J.; Xu, S.; Gao, S.; Liu, H.; Qiu, H.; Li, Z.; Liu, H.; Li, Z.; Li, H. Growth Graphene on Silver− Copper Nanoparticles by Chemical Vapor Deposition for HighPerformance Surface-Enhanced Raman Scattering. Appl. Surf. Sci. 2015, 353, 63−70. (24) Gellini, C.; Sabatino, G.; Papini, A. M.; Muniz-Miranda, M. SERS Study of a Tetrapeptide Based on Histidine and Glycine Residues, Adsorbed on Copper/Silver Colloidal Nanoparticles. J. Raman Spectrosc. 2014, 45, 418−423. (25) Chen, X.; Cui, C.; Guo, Z.; Liu, J.; Huang, X.; Yu, S. Unique Heterogeneous Silver−Copper Dendrites with a Trace Amount of Uniformly Distributed Elemental Cu and Their Enhanced SERS Properties. Small 2011, 7, 858−863. (26) Li, D.; Liu, J.; Wang, H.; Barrow, C. J.; Yang, W. Electrochemical Synthesis of Fractal Bimetallic Cu/Ag Nanodendrites for Efficient Surface Enhanced Raman Spectroscopy. Chem. Commun. 2016, 52, 10968−10971. (27) Chen, L. Y.; Zhang, L.; Fujita, T.; Chen, M. W. SurfaceEnhanced Raman Scattering of Silver@Nanoporous Copper CoreShell Composites Synthesized by an In Situ Sacrificial Template Approach. J. Phys. Chem. C 2009, 113, 14195−14199. (28) Porel, S.; Singh, S.; Harsha, S. S.; Rao, D. N.; Radhakrishnan, T. P. Nanoparticle-Embedded Polymer: In Situ Synthesis, Free-standing Films with Highly Monodisperse Silver Nanoparticles and Optical Limiting. Chem. Mater. 2005, 17, 9−12. (29) Ramesh, G. V.; Porel, S.; Radhakrishnan, T. P. Polymer Thin Films Embedded with In Situ Grown Metal Nanoparticles. Chem. Soc. Rev. 2009, 38, 2646−2656. (30) Hariprasad, E.; Radhakrishnan, T. P. Chemistry Inside a Polymer Thin Film: In Situ Soft Chemical Synthesis of Metal Nanoparticles and Applications. In Nanocomposites: In Situ Synthesis of Polymer-Embedded Nanostructures; Carotenuto, G., Nicolais, L., Eds.; John Wiley and Sons: Hoboken, NJ, 2013; pp129−144. (31) Rao, V. K.; Radhakrishnan, T. P. Hollow Bimetallic Nanoparticles Generated In Situ Inside a Polymer Thin Film: Fabrication and Catalytic Application of Silver−Palladium−Poly(vinyl alcohol). J. Mater. Chem. A 2013, 1, 13612−13618. (32) Rao, V. K.; Radhakrishnan, T. P. Tuning the SERS Response with Ag-Au Nanoparticle − Embedded Polymer Thin Film Substrates. ACS Appl. Mater. Interfaces 2015, 7, 12767−12773. (33) Hariprasad, E.; Radhakrishnan, T. P. In Situ Fabricated Polymer−Silver Nanocomposite Thin Film as an Inexpensive and Efficient Substrate for Surface-Enhanced Raman Scattering. Langmuir 2013, 29, 13050−13057. (34) Jaffe, H. L.; Rosenblum, F. M. Poly(vinyl alcohol) for Adhesives. In Handbook of Adhesives; Skeist, I., Ed.; Chapman and Hall: New York, 1990; pp 401−407. (35) Gawande, M. B.; Goswami, A.; Felpin, F.; Asefa, T.; Huang, X.; Silva, R.; Zou, X.; Zboril, R.; Varma, R. S. Cu and Cu-Based Nanoparticles: Synthesis and Applications in Catalysis. Chem. Rev. 2016, 116, 3722−3811. (36) Rehman, S.; Mumtaz, A.; Hasanain, S. K. Size Effects on the Magnetic and Optical Properties of CuO Nanoparticles. J. Nanopart. Res. 2011, 13, 2497−2507. (37) Dagher, S.; Haik, Y.; Ayesh, A. I.; Tit, N. Synthesis and Optical Properties of Colloidal CuO Nanoparticles. J. Lumin. 2014, 151, 149− 154. (38) D'Andrea, C.; Irrera, A.; Fazio, B.; Foti, A.; Messina, E.; Maragò, O. M.; Kessentini, S.; Artoni, P.; David, C.; Gucciardi, P. G. Red Shifted Spectral Dependence of the SERS Enhancement in a Random Array of Gold Nanoparticles Covered with a Silica Shell: Extinction versus Scattering. J. Opt. 2015, 17, 114016. I

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C (39) El-Safty, S. A.; Ismail, A. A.; Matsunaga, H.; Nanjo, H.; Mizukami, F. Uniformly Mesocaged Cubic Fd3m Monoliths as Model Carriers for Optical Chemosensors. J. Phys. Chem. C 2008, 112, 4825− 4835. (40) Christian, G. D. Analytical Chemistry, 6th ed.; John Wiley and Sons: New York, 2003.

J

DOI: 10.1021/acs.jpcc.6b10238 J. Phys. Chem. C XXXX, XXX, XXX−XXX